Abstract
Mitochondrial DNA (mtDNA) mutations may contribute to aging and age-related diseases. Previously, we reported that accumulation of mtDNA mutations is associated with age-related hearing loss in mice carrying a mutator allele of the mitochondrial Polg DNA polymerase. To elucidate the role of mtDNA mutations in the pathogenesis of age-related hearing loss or presbycusis, we performed large scale gene expression analysis to identify mtDNA mutation-responsive genes and biological process categories associated with mtDNA mutations by comparing the gene expression patterns of cochlear tissues from 9-month-old mitochondrial mutator and control mice. mtDNA mutations were associated with transcriptional alterations consistent with impairment of energy metabolism, induction of apoptosis, cytoskeletal dysfunction, and hearing dysfunction in the cochlea of aged mitochondrial mutator mice. TUNEL staining and caspase-3 immunostaining analysis demonstrated that the levels of apoptotic markers were significantly increased in the cochleae of mitochondrial mutator mice compared to age-matched controls. These observations support a new model of how mtDNA mutations impact cochlear function whereby accumulation of mtDNA mutations lead to mitochondrial dysfunction, an associated impairment of energy metabolism, and the induction of an apoptotic program. The data presented here provide the first global assessment at the molecular level of the pathogenesis of age-related disease in mitochondrial mutator mice and reveal previously unrecognized biological pathways associated with mtDNA mutations.
Keywords: mtDNA mutations, Polg, Age-related hearing loss, Presbycusis, Apoptosis, Mitochondrial dysfunction, Aging
1. Introduction
The mitochondrial theory of aging postulates that reactive oxygen species (ROS) generated inside mitochondria damage key mitochondrial components such as mtDNA, membranes, and respiratory chain proteins (Balaban et al., 2005). Such damage accumulates with time, leads to age-related mitochondrial dysfunction, and causes energy depletion, compromising the organism. Mitochondria are the main source of cellular energy, generating most cellular ATP. The mammalian mitochondrial genome consists of 37 genes, encoding 13 proteins of the electron transport oxidative phosphorylation system (Anderson et al., 1981). Since mtDNA plays an essential role in energy metabolism, mtDNA mutations have been hypothesized to contribute to mammalian aging. Indeed, previous studies have shown that mtDNA point mutations accumulate with aging in humans (Michikawa et al., 1999), and that accumulation of mtDNA mutations leads to premature aging in mitochondrial mutator mice, indicating a causal role of mtDNA mutations in mammalian aging (Kujoth et al., 2005; Trifunovic et al., 2004).
Specific mutations in mtDNA have also been implicated in the development of age-related diseases (Corral-Debrinski et al., 1992; Dirks et al., 2006; Seidman et al., 2002). Age-related hearing loss, also known as presbycusis, is one of the most common diseases of aging and the number of people affected is rapidly growing as the elderly population increases. However, the pathogenesis of age-related hearing loss remains unclear. The accumulation of mtDNA mutations has been associated with the development of age-related hearing loss (Kujoth et al., 2005). In support of this hypothesis, hearing loss is a common symptom in individuals harboring inherited mtDNA mutations (Chinnery et al., 2000). Several mutations in Polg (mitochondrial DNA polymerase gamma) have been identified as a cause of human disorders such as Alpers syndrome and deafness (Mancuso et al., 2004; Nguyen et al., 2005). Specific mtDNA point mutations also contribute to mitochondrial disorders in humans such as mitochondrial encephalomyopathy, lactic acidosis and stroke-like episodes (MELAS) and myoclonic epilepsy and ragged red fibres (MERRF), the symptoms of which include hearing loss (Chinnery et al., 2000; Fischel-Ghodsian, 2003). Furthermore, more than 100 different deletions of mtDNA have been associated with mitochondrial disorders (Chinnery et al., 2000), and of these, some mtDNA deletions cause Kearns–Sayre syndrome which also involves hearing loss (Chinnery et al., 2000).
We have shown previously that mtDNA mutations play a role in the development of age-related hearing loss in mice carrying a mutation in Polg (Kujoth et al., 2005). These mice were created by introducing a two base substitution, which results in a critical residue substitution in the exonuclease domain of Polg, impairing its mtDNA proof-reading ability (Kujoth et al., 2005). These animals accumulate mtDNA mutations with aging, and exhibit a striking premature aging phenotype, consisting of osteoporosis, reduced lifespan, and hearing loss (Kujoth et al., 2005; Trifunovic et al., 2004). Thus, these mice represent a useful model for investigating a role of mtDNA mutations in the pathogenesis of age-related hearing loss. In the present study, we performed genome-wide DNA microarray analysis using cochlear tissues from 9-month-old mitochondrial mutator and control mice. We then performed data analysis to identify age-related, mtDNA mutation-responsive genes and biological process categories statistically associated with mtDNA mutation-responsive genes. Quantitative RT-PCR (QRT-PCR), TUNEL staining, immunostaining, histology, and hearing assessment were performed to corroborate the microarray results. Together, these approaches revealed previously unrecognized biological process pathways associated with mtDNA mutations, providing a new model of how mtDNA mutations impact cochlear function.
2. Materials and methods
2.1. Animals
PolgD257A/D257A mice have been previously described and were backcrossed four generations onto the C57BL/6 background (Kujoth et al., 2005). Mice were housed in the Genetics and Biotechnology Center of University of Wisconsin-Madison-approved Animal Care Facility. Experiments were performed in accordance with the protocols approved by the University of Wisconsin Institutional Animal Care and Use Committee (Madison, WI).
2.2. Assessment of hearing
Detailed protocols for ABR measurements have been described (Someya et al., 2007). Briefly, ABRs were measured with a tone burst stimulus (4, 8, and 16 kHz) using an ABR recording system (Intelligent Hearing System, Miami, FL). Mice were anesthetized with a mixture of xylazine hydrochloride (10 mg/kg, i.m.) and ketamine hydrochloride (40 mg/kg, i.m.). We used five mice per group for this study. Following the hearing measurements, the same mice were sacrificed to conduct microarray analysis of cochlear gene expression. Data were analyzed using the two-tailed standard t-test. All data were reported as mean ± S.E.M.
2.3. Assessment of histology
Detailed protocols for tissue processing have been described (Sakamoto et al., 2004). Briefly, 4% paraformaldehyde-fixed and paraffin-embedded specimens were sliced into 4µm sections, mounted on silane-coated slides, stained with Haematoxylin and Eosin (HE), and observed under a light microscope (Leica Microsystems, Bannockburn, IL). The Rosenthal’s canal was divided into three regions: apical, middle and basal (Keithley et al., 2004) and the apical and basal turns were used for evaluation of cochlear histology. We used five mice for each group (n = 5) that were different mice than those used in the microarray analysis. Five modiolar sections obtained in every fifth section from one unilateral cochlea were evaluated per mouse. The same animals were used for spiral ganglion cell counting, TUNEL staining, and immunostaining.
Spiral ganglion cells (SGCs) were counted in the apical turn of the cochlear sections in the field of 0.3 µm ± 0.225 µm as seen using a 40 × objective by direct observation. SGCs were identified by the presence of a nucleus. The SGC density was calculated as the number of SGCs per 1mm2. Five sections of the unilateral apical turn were evaluated in one cochlea per mouse. Data were analyzed using the two-tailed standard t-test. All data were reported as mean ± S.E.M.
2.4. RNA sample preparations
Detailed protocols for gene expression profiling analysis using Affymetrix microarrays by Affymetrix algorithms have been described (Affymetrix, 2004; Lee et al., 1999). Briefly, we pooled two cochleae from each mouse for one sample because the amount of mRNA from one cochlea of one mouse was insufficient for a single GeneChip (Affymetrix, Santa Clara, CA).We hybridized each sample to a single GeneChip and used five samples per group.
2.5. Statistical analysis
Signals in each image were scaled to a whole chip target intensity of 1500 using GeneChip® Operating Software (GCOS) 1.3 in order to minimize an overall variability in hybridization intensities. A gene probe set was considered “expressed” if it displayed a “present” call in at least one GeneChip based on the Affymetrix “present/absent call” algorithms. All genes considered “not expressed” were eliminated from our analysis. Further analysis and comparisons between the two groups were also done using GCOS 1.3 and Significance Analysis of Microarray (SAM) (Tusher et al., 2001). To identify mtDNA-responsive genes, each 9-monthold WT sample (n = 5) was compared to each 9-month-old D257A sample (n = 5), generating a total of 25 pairwise comparisons. The change in gene expression was considered significant if the fold change was >1.0 based on the Affymetrix algorithms (Affymetrix, 2004), the P-value was <0.02 using the two-tailed standard t-test, and FDR was <10% based on the SAM algorithms (Tusher et al., 2001). We then used Database for Annotation, Visualization, and Integrated Discovery (DAVID) (Dennis et al., 2003) and Expression Analysis Systematic Explorer (EASE) (Hosack et al., 2003) (http://apps1.niaid.nih.gov/david/) to assign identified genes to “Gene Ontology (GO): Biological Process” categories of Gene Ontology Consortium (www.geneontology.org) and used EASE to determine the total number of identified genes that were assigned to each biological process category, the total number of genes on the array in each biological process category, and to identify “GO: Biological Process” categories statistically associated with mtDNA mutation-responsive genes by performing Fisher Exact tests (Fisher Exact score < 0.05). Five genes were manually assigned to “perception of sound” category using the gene list of “Hereditary Hearing Impairment in Mice” by The Jackson Laboratory (http://www.jax.org/hmr/master_table.html). The complete set of microarray data has been submitted to the GEO (Gene Expression Omnibus) repository (http://www.ncbi.nlm.nih.gov/geo/) with GEO Accession number GSE4866. For further details of the methods (MIAME checklist), see Supplementary information (Supplementary note).
2.6. Quantitative RT-PCR
Detailed protocols for QRT-PCR analysis have been described (Someya et al., in press). Detection of mRNA was carried out with the TaqMan EZ RT-PCR kit using the Applied Biosystems Prism 7000 Sequence Detection System (Applied Biosystems, Foster City, CA). Oligonucleotide primers and MGB fluorescent probes (TaqMan Gene Expression Assays) were purchased from Applied Biosystems. Each sample used for QRT-PCR analysis consisted of the two cochleae pooled from one mouse. We used four mice per group (n = 4) for this study. mRNA samples used for QRT-PCR analysis were from different mice than those used for microarray analysis which had been exhausted for that purpose. All data were reported as mean ± S.E.M.
2.7. TUNEL staining
Detailed protocols for TUNEL staining have been described (Someya et al., 2007). Briefly, TUNEL staining for apoptotic nuclei was performed using the DeadEnd Colorimetric TUNEL System (Promega, Madison, WI) according to the manufacturer’s instructions. Briefly, stained sections were viewed by light microscopy and spiral ganglions in the apical turns were evaluated for TUNEL positivity. We used five mice for each group (n = 5) that were different mice than those used in the microarray analysis, but the same as those used for spiral ganglion cell counting and immunostaining. Data were evaluated using the two-tailed standard t-test. All data were reported as mean ± S.E.M.
2.8. Immunostaining
Detailed protocols for immunohistochemical staining have been described (Someya et al., 2007). Immunohistochemical staining for caspase-3 was performed using the Anti-Active Caspase-3 pAb (Promega, Madison, WI) according to the manufacturer’s instructions. Briefly, positively-stained cells were observed in sections of the cochleae by direct observation using a light microscopy. Spiral ganglions in the apical turns were evaluated for caspase-3 positivity. We used five mice for each group (n = 5) that were different mice than those used in the microarray analysis, but the same as those used for spiral ganglion cell counting and TUNEL staining. Data were evaluated using the two-tailed standard t-test. All data were reported as mean ± S.E.M.
3. Results
3.1. Assessment of hearing and histology
The C57BL/6 strain possesses mutations in the Cdh23 gene which result in onset of age-related hearing loss and cochlear degeneration beginning at 8–10 months (Keithley et al., 2004; Noben-Trauth et al., 2003). Because the background of our mice is C57BL/6 and both PolgD257A/D257A (D257A) and Polg+/+ wild-type (WT) mice possess the mutations, we first evaluated hearing function of 8–9-month-old WT mice and confirmed that they did not show any age-related hearing loss at the frequencies examined under the housing conditions of our animal facility. We then chose to study 9-month-old WT mice and age-matched D257A mice, an age at which WT mice do not have hearing loss due to the Cdh23 mutations.
To determine whether D257A mice display age-related hearing loss, we measured auditory brainstem response (ABR) thresholds in 9-month-old WT and D257A mice. Nine-month-old WT mice displayed normal hearing. However, the age-matched D257A mice displayed significant age-related hearing loss (P < 0.05, n=5) (Fig. 1). To confirm whether D257A mice display age-related cochlear degeneration, we performed histological examination of the cochleae from 2- and 9-month-old WT and D257A mice (n = 5). Examination of the histology of the apical cochlear region confirmed that two of the five cochleae from 9-month-old D257A mice displayed severe loss of SGCs (Fig. 2J and L). Consistent with our previous studies, all the cochleae from 9-month-old D257A mice displayed severe loss of SGCs in the basal turn (Fig. 3G) (Kujoth et al., 2005). In contrast, all the cochleae from 2-month-old WT and D257A mice displayed no or only a few loss of SGCs (Fig. 2A, C, D, and F). The five cochleae from 9-month-old WT mice also displayed no clear loss of SGCs (Fig. 2G and I). Decreased SGC density is one of the hallmarks of presbycusis in mammals (Keithley et al., 2004). The mean SGC density of 9-month-old D257A mice was significantly lower than that of age-matched WT (Fig. 2M) (P < 0.05, n = 5). There was no significant difference in the SGC density between 2-monthold WT, age-matched D257A, and 9-month-old WT mice (Fig. 2M). Hair cell loss is another hallmark of presbycusis (Francis et al., 2003). Examination of the histology of the basal cochlear region confirmed that two of the five cochleae from 9-month-old D257A mice displayed severe loss of outer hail cells (OHCs) and inner hair cells (IHCs) (Fig. 3G and H), whereas all the cochleae from 2-month-old WT and D257A and 9-month-old WT mice displayed no or only a few loss of the OHCs and IHCs (Fig. 3A–F). All the cochleae from 2- and 9-month-old WT and D257A mice also displayed no or only a few loss of the OHCs and IHCs in the apical cochlear region (Fig. 2B, E, H, and K). Together, these observations confirm that D257A mice display age-related hearing loss and cochlear degenerations by 9 months of age.
Fig. 1.
Age-related hearing loss. Means of ABR thresholds (dB SPL, decibels sound pressure level) for 9-month-old WT and D257A mice at 4, 8, and 16 kHz. The mean ABR thresholds of 9-month-old D257A mice were significantly elevated compared to the age-matched WT mice (*P < 0.05, n = 5), showing age-related hearing loss. Error bars represent S.E.M.
Fig. 2.
Age-related cochlear degeneration. Cochlear apical turn from 2-month-old WT (A–C) and D257A (D–F), and 9-month-old WT (G–I) and D257A (J–L) mice. Arrows indicate hair cells and spiral ganglion cells (SGCs). The cochleae from 2-month-old WT and D257A, and 9-month-old WT mice displayed no or only a few loss of the hair cells and spiral SGCs, whereas the cochleae from 9-month-old D257A mice displayed severe loss of SGCs. The mean SGC density of D257A mice was significantly lower than that of age-matched WT (M) (*P < 0.05, n = 5). Error bars represent S.E.M. Scale bar = 100µm (left panel). Scale bar = 25µm (middle and right panel).
Fig. 3.
Age-related hair cell loss. Cochlear basal turn from 2-month-oldWT (A and B) and D257A (C and D), and 9-month-oldWT(E and F) and D257A (G and H) mice. Arrows indicate hair cells. The cochleae from 9-month-old D257A mice displayed severe loss of the out hair cells (OHCs) and inner hair cells (IHCs), whereas the cochleae from 2-month-old WT and D257A, and 9-month-old WT mice displayed no or only a few loss of the OHCs and IHCs. Scale bar = 100µm (left panel). Scale bar = 25µm (right panel).
3.2. Overview of microarray and statistical analysis
To identify mtDNA mutation-responsive genes and associated biological process categories, we conducted genome-wide gene expression analysis using RNA samples isolated from cochlear tissues of 9-month-old WT and D257A mice (n = 5). We selected cochlear tissue for our investigation because the cochlea is primarily composed of long-lived, high oxygen-consuming postmitotic cells (Pickles, 2004), a feature shared with other critical postmitotic tissues such as heart and brain which are sites of multiple, major age-related diseases and show accumulation of mtDNA mutations (Corral-Debrinski et al., 1992). We used the Affymetrix mouse genome 430 2.0 array, which contains 45,037 gene probe sets representing more than 34,000 mouse genes and ESTs. We then performed statistical analysis using the Affymetrix algorithms (Affymetrix, 2004), SAM algorithms (Tusher et al., 2001), and our own data analysis criteria. We found that 1289 gene probe sets were significantly up-regulated, and 662 gene probe sets were significantly down-regulated in cochlear tissues from 9-month-old D257A mice. The gene probe sets were further assigned to “GO: Biological Process” categories of the Gene Ontology Consortium using DAVID (Dennis et al., 2003), resulting in classification of 466 up-regulated and 378 down-regulated genes. Gene probe sets were considered “genes” if they had been assigned a “gene symbol” annotation by DAVID. We then used EASE (Hosack et al., 2003) to determine the number of identified genes and the total number of genes on the array in each biological process category, and to test the statistical significance of coregulation (overrepresentation) of identified genes within each biological process category. For this analysis, we used the Fisher Exact test (P < 0.05), whose score represents the probability that an overrepresentation of a certain biological process category occurs by random chance (Hosack et al., 2003). A summary of the Gene Ontology Biological Process categories statistically associated with mtDNA-responsive genes is shown in Table 1. A list of selected genes altered in expression is shown in Table 2. For all genes identified, see Supplementary information (Table 3).
Table 1.
List of “GO: Biological Process” categories of genes significantly associated with mtDNA mutation-responsive genes in the aging cochlea
| Biological process categories | N | TN | Fisher |
|---|---|---|---|
| Up-regulated (466 classified genes) | |||
| Ribonucleoprotein complex | 23 | 242 | 0.000 |
| Ribosome | 14 | 147 | 0.003 |
| RNA binding | 21 | 285 | 0.005 |
| Protein binding | 71 | 1401 | 0.020 |
| Anti-apoptosis | 5 | 41 | 0.023 |
| Regulation of apoptosis | 10 | 123 | 0.025 |
| Intracellular transport | 25 | 423 | 0.031 |
| Down-regulated (378 classified genes) | |||
| Cytoskeleton | 39 | 440 | 0.000 |
| Extracellular matrix | 25 | 237 | 0.000 |
| Collagen | 8 | 32 | 0.000 |
| Structural molecule activity | 32 | 456 | 0.000 |
| Cell motility | 15 | 158 | 0.000 |
| Calcium ion binding | 28 | 421 | 0.000 |
| Myosin | 6 | 33 | 0.001 |
| Cell adhesion | 22 | 425 | 0.017 |
| Carbohydrate metabolism | 13 | 239 | 0.044 |
N: the number of identified genes in the category; FC: fold change; TN: the total number of genes in the category represented on the GeneChip; Fisher: Fisher Exact test score.
Table 2.
List of selected genes significantly associated with mtDNA mutations
| Gene ID | FC | P-value | FDR | Symbol |
|---|---|---|---|---|
| Regulation of apoptosis (10) | ||||
| BB831420 | 1.69 | 0.000 | 2.9 | Gsk3b |
| AA214868 | 2.07 | 0.000 | 3.2 | Pten |
| BM120925 | 1.66 | 0.000 | 4.0 | Bim |
| NM 007566 | 1.37 | 0.001 | 4.1 | Birc6 |
| AF201288 | 1.11 | 0.003 | 6.2 | Tsc22d3 |
| BB768208 | 1.25 | 0.004 | 6.6 | Sgk3 |
| AB012278 | 1.35 | 0.006 | 7.5 | Cebpb |
| NM 026810 | 1.23 | 0.009 | 8.9 | Mlh1 |
| AK011828 | 1.66 | 0.010 | 8.9 | Hip1 |
| AF353993 | 1.32 | 0.013 | 9.7 | Bbp |
| Programmed cell death (13) | ||||
| NM 011361 | 1.56 | 0.000 | 3.8 | Sgk |
| BC026823 | 1.39 | 0.002 | 5.6 | Pdcd6ip |
| L26349 | 1.26 | 0.003 | 5.9 | Tnfrsf1a |
| Proteolysis and peptidolysis (14) | ||||
| BB443585 | 2.37 | 0.002 | 5.4 | Adamts4 |
| AK008695 | 1.63 | 0.004 | 6.2 | 2210010C04Rik |
| NM 022022 | 1.08 | 0.005 | 6.6 | Ube4b |
| AK020649 | 1.45 | 0.005 | 6.9 | Adam 15 |
| BB131619 | 1.62 | 0.006 | 7.5 | Lnx1 |
| NM 007800 | 2.27 | 0.007 | 8.2 | Ctsg |
| AK013777 | 2.42 | 0.008 | 8.2 | Ptpn21 |
| NM 008585 | 1.55 | 0.008 | 8.2 | Mep1a |
| NM 133216 | 1.28 | 0.011 | 8.9 | Xpnpep1 |
| BC013654 | 1.56 | 0.011 | 8.9 | Trfr2 |
| Response to stress (15) | ||||
| AF047377 | 1.42 | 0.002 | 5.0 | Hspb1 |
| BF466143 | 2.23 | 0.003 | 5.9 | Wasl |
| BM234360 | 1.54 | 0.003 | 5.9 | Fn1 |
| L77884 | 2.62 | 0.003 | 5.9 | Lyst |
| NM 011558 | 2.27 | 0.004 | 6.6 | Tcrg-V4 |
| BB298208 | 3.39 | 0.005 | 6.6 | Chek1 |
| BB075261 | 1.48 | 0.005 | 6.9 | Fcgr1 |
| NM 016928 | 1.64 | 0.006 | 7.5 | Tlr5 |
| NM 011728 | 1.46 | 0.007 | 8.2 | Xpa |
| NM 010824 | 2.01 | 0.011 | 8.9 | Mpo |
| Carbohydrate metabolism (13) | ||||
| BQ176370 | −1.47 | 0.00 | 2.9 | 9230112O05Rik |
| NM 008063 | −1.29 | 0.000 | 4.0 | Slc37a4 |
| NM 008832 | −1.83 | 0.001 | 5.0 | Phka1 |
| AW208566 | −4.83 | 0.001 | 5.0 | Lyzs |
| BG073164 | −1.42 | 0.002 | 6.2 | Pgm2l1 |
| NM 007933 | −5.77 | 0.003 | 6.9 | Eno3 |
| BI685536 | −1.35 | 0.004 | 7.5 | B4galt6 |
| NM 008062 | −1.34 | 0.004 | 7.5 | G6pdx |
| AK007262 | −1.57 | 0.004 | 8.2 | 1700124B08Rik |
| NM 011079 | −3.91 | 0.004 | 8.2 | Phkg |
| Mitochondrion (13) | ||||
| BC008518 | −1.16 | 0.000 | 3.8 | Nnt |
| NM 007702 | −2.69 | 0.000 | 4.0 | Cidea |
| AB022340 | −4.11 | 0.001 | 4.5 | Acsm3 |
| AV278574 | −1.42 | 0.001 | 5.4 | Aldh6a1 |
| NM 007472 | −1.65 | 0.002 | 6.2 | Aqp1 |
| NM 080575 | −1.57 | 0.003 | 6.9 | Acas2l |
| NM 026331 | −1.89 | 0.004 | 7.5 | Mscp |
| NM 022032 | −1.49 | 0.004 | 7.5 | Perp |
| AK017491 | −1.47 | 0.004 | 8.2 | Mipep |
| BG866501 | −1.38 | 0.005 | 8.9 | Hadhb |
| Cytoskeleton (39) | ||||
| BG094386 | −1.27 | 0.000 | 2.9 | Dag1 |
| BQ032637 | −1.49 | 0.000 | 3.8 | Jak1 |
| BC021484 | −1.70 | 0.000 | 3.8 | Sspn |
| AK003186 | −1.98 | 0.000 | 4.0 | Tpm2 |
| AF237628 | −38.21 | 0.000 | 4.0 | Lmod2 |
| AK003182 | −7.87 | 0.000 | 4.0 | Myl1 |
| AJ002522 | −8.77 | 0.001 | 5.0 | My h1 |
| NM 010861 | −3.49 | 0.002 | 6.2 | Myl2 |
| BC008538 | −136.05 | 0.003 | 6.2 | Myh2 |
| NM 016754 | −6.87 | 0.005 | 8.2 | Mylpf |
| Extracellular matrix (25) | ||||
| AW550625 | −1.91 | 0.000 | 2.9 | Col3a1 |
| BF227507 | −2.95 | 0.000 | 2.9 | Col1a2 |
| NM 008610 | −1.74 | 0.000 | 3.8 | Mmp2 |
| BC019502 | −1.71 | 0.000 | 3.8 | Bgn |
| AV229424 | −1.58 | 0.000 | 4.0 | Col5a2 |
| BF158638 | −1.35 | 0.001 | 5.4 | Col4a1 |
| NM 007732 | −14.01 | 0.002 | 5.6 | Col1 7a 1 |
| AF011450 | −1.30 | 0.003 | 6.6 | Col15a1 |
| AF064749 | −1.33 | 0.003 | 6.6 | Col6a3 |
| U08020 | −3.18 | 0.004 | 7.5 | Col1a1 |
| Cell adhesion (22) | ||||
| BE199556 | −1.43 | 0.000 | 3.8 | Pard3 |
| AF038562 | −5.52 | 0.000 | 3.8 | Lgals7 |
| NM 009675 | −2.53 | 0.000 | 3.8 | Aoc3 |
| BI110565 | −2.01 | 0.000 | 4.0 | Postn |
| BB776961 | −1.33 | 0.001 | 4.5 | Cdh13 |
| NM 019759 | −2.14 | 0.002 | 5.6 | Dpt |
| BB534670 | −2.31 | 0.003 | 6.9 | Cd36 |
| NM 11582 | −2.92 | 0.004 | 6.9 | Thbs4 |
| BQ175493 | −1.56 | 0.004 | 7.5 | Itga8 |
| NM 016685 | −2.35 | 0.007 | 9.7 | Comp |
| Perception of sound (6) | ||||
| BI143915 | −1.69 | 0.001 | 4.3 | Ntn1* |
| BQ176524 | −1.72 | 0.001 | 5.4 | Bsn* |
| U08020 | −3.18 | 0.004 | 7.5 | Col1a1* |
| BQ175493 | −1.56 | 0.004 | 7.5 | Itga8* |
| NM 010081 | −6.85 | 0.006 | 8.9 | Dst* |
| NM 013624 | −1.52 | 0.007 | 9.7 | Otog |
| Ion transport (12) | ||||
| BG791642 | −1.73 | 0.001 | 4.4 | Abcc9 |
| BE945884 | −2.61 | 0.001 | 5.0 | Gabra1 |
| BQ175666 | −1.55 | 0.001 | 5.4 | Gabrb3 |
| NM 013540 | −2.04 | 0.004 | 6.9 | Gria2 |
| BB515151 | −1.39 | 0.005 | 8.2 | Tcn2 |
| BM119753 | −1.51 | 0.006 | 9.7 | Kcna6 |
| AF089751 | −1.49 | 0.006 | 9.7 | P2rx4 |
| BB236747 | −1.48 | 0.006 | 9.7 | Clic5 |
| BI650739 | −1.33 | 0.007 | 9.7 | Ubl4 |
| NM 08557 | −1.57 | 0.007 | 9.7 | Fxyd3 |
Gene ID: representative public gene ID; FC: fold change; P-value: P-values for the comparison of 9-month-oldWTvs. 9-month-old D257A group; FDR: false discovery rate; symbol: gene symbol.
3.3. mtDNA mutations result in up-regulation of genes involved in apoptosis and stress response
EASE analysis revealed that two apoptotic process categories, “regulation of apoptosis” and “anti-apoptosis”, were significantly associated with the mtDNA mutation-responsive genes (Fisher Exact score < 0.05), and 15 genes within these categories were significantly up-regulated in the cochleae (P < 0.02, false discover rate (FDR) < 10%) (Table 1). Of the 15 genes identified, 10 genes were involved in regulation of apoptosis, including phosphatase and tensin homolog (Pten), BCL2-like 11 (apoptosis facilitator) (Bim), baculoviral IAP repeat-containing 6 (Birc6), and beta-amyloid binding protein precursor (Bbp) (Table 2). Bim is a member of the proapoptotic Bcl2 family and a well-described component of mitochondrial apoptosis pathway, leading to cytochrome c release (Cheng et al., 2001; Putcha et al., 2001). Birc6 is also involved in mitochondrial pathways of apoptosis (Ren et al., 2005). Phosphatase and tensin homolog is a tumor suppressor phosphatase that regulates Akt kinase activity and may promote apoptotic death of hippocampal neurons (Gary and Mattson, 2002). The EASE analysis also revealed that 13 genes in the “programmed cell death” category were up-regulated, including serum/glucocorticoid regulated kinase (Sgk), programmed cell death 6 interacting protein (Pdcd6ip), and tumor necrosis factor receptor superfamily, member 1a (Tnfrsf1a) (P < 0.02, FDR < 10%, Fisher Exact score = 0.21) (Table 2). Sgk is a member of a serine/threonine kinase family and is involved in the early pathogenesis of Parkinson’s disease (Schoenebeck et al., 2005). The protein encoded by Pdcd6ip regulates neuronal death involving caspase pathways (Trioulier et al., 2004). Tnfrsf1a is a member of a TNFR superfamily and is involved in activation of Bid, a Bcl-2 family member and inducer of mitochondrial pathways of apoptosis (Yin, 2000). QRT-PCR analysis was conducted for Pten and Bim, to validate the microarray results for expression of these genes. The QRT-PCR results were in agreement with the microarray findings that expression of these proapoptotic genes were increased in the cochlea of D257A mice (Fig. 4).
Fig. 4.
QRT-PCR validation of microarray data. The data represent the fold-change in gene expression of 9-month-old D257A cochlea compared to age-matched WT based on average signal intensities of microarray hybridizations (n = 4) and gene expression levels determined by quantitative RT-PCR (n = 4). A consistent agreement between fold-change determined by the two methods was observed. Error bars represent SEM. Fold change: 9-month-old WT to age-matched D257A ratio. QRT-PCR: quantitative RTPCR.
Proteolysis also plays a key role in cell death (Hengartner, 2000). The EASE analysis revealed that 14 genes involved in the “proteolysis and peptidolysis” category were found to be up-regulated, including a disintegrin-like and metalloprotease (reprolysin type) with thrombospondin type 1 motif, 4 (Adamts4), a disintegrin and metalloproteinase domain 15 (metargidin) (Adam15), and cathepsin G (Ctsg) (P < 0.02, FDR < 10%, Fisher Exact score = 0.88) (Table 2). The proteins encoded by Adamts4 and Adam15 are involved in proteolysis of neurocan, aggrecan, and versican, members of a proteoglycan family, and components of the extracellular matrix in various tissues (Gao et al., 2002). Cathepsins are cysteine proteases that are released from lysosomes and play a key role in apoptotic cell death (Chwieralski et al., 2006). Up-regulation of these proteolysis genes provides further evidence of increased transcriptional apoptotic activity in the aging cochlea (Table 2).
Stress responses may trigger apoptotic cell death (Mattson, 2000). The EASE analysis revealed that 15 genes involved in the “response to stress” category were up-regulated, including heat shock protein 1 (Hspb1) and myeloperoxidase (Mpo) (P < 0.02, FDR < 10%, Fisher Exact score = 0.65) (Table 2). The generation of ROS plays an important role in immune-mediated defense against infection, tumor cells, and cell damage (Aratani et al., 1999). Myeloperoxidase produces hypochlorous acid from hydrogen peroxide and chloride, and plays a key role in microbial killing or immune-mediated cell destruction (Aratani et al., 1999). In the mammalian brain, small heat shock proteins are induced in response to physiological stress such as tissue injury (Murashov et al., 1998). Hspb1 is a member of the small heat shock protein family and has been characterized as a molecular chaperone required for protection of motor neurons under conditions of stress (Murashov et al., 1998). Together, these results provide evidence that mtDNA mutations result in significant up-regulation of genes involved in multiple apoptotic and stress pathways in the cochlea of aged mitochondrial mutator mice.
3.4. mtDNA mutations result in down-regulation of genes involved in energy metabolism
The EASE analysis revealed that one of the key energy metabolism categories, “carbohydrate metabolism”, was significantly suppressed in D257A mice (Fisher Exact score < 0.05), and that 13 genes in this category were significantly down-regulated in the cochlea (P < 0.02, FDR < 10%) (Table 1). These include phosphorylase kinase alpha 1 (Phka1), phosphoglucomutase 2-like 1 (Pgm2l1), and phosphorylase kinase gamma 1 (Phkg1) (Table 2). The glycogen phosphorylase kinases such as phosphorylase kinase alpha 1 and phosphorylase kinase gamma 1 are key enzymes involved in glycogen metabolism and regulate glycogen breakdown by activating glycogen phosphorylase (Wuyts et al., 2005). Pgm2l1 is a member of a Phosphoglucomutase gene family involved in glycolysis and gluconeogenesis (Levin et al., 1999). Phosphoglucomutases regulate the interconversion of glucose 1-phosphate and glucose 6-phosphate in the glycolysis pathway (Levin et al., 1999). The EASE analysis also revealed that 13 genes in the “mitochondrion” category were down-regulated, including aquaporin 1 (Aqp1), PERP, TP53 apoptosis effector (Perp), and hydroxyacyl-Coenzyme A dehydrogenase/3-ketoacyl-Coenzyme A thiolase/enoyl-Coenzyme A hydratase (trifunctional protein), beta subunit (Hadhb) (P < 0.02, FDR < 10%, Fisher Exact score = 0.79) (Table 2). Aquaporins are membrane proteins that regulate water transport. Specific members of this water channel family have been shown to be expressed in the mitochondria of mammalian brain (Amiry-Moghaddam et al., 2005). Aquaporin 1, a member of the water channel family is localized in the stria vascularis of rat cochlea and is thought to play a key role in ion transport of the auditory system (Sawada et al., 2003). The protein encoded by Hadhb is a subunit of a multienzyme complex or mitochondrial functional protein that is involved in mitochondrial µ-oxidation (Orii et al., 1997). QRT-PCR analysis was conducted for Eno3 and Acsm3 from these energy metabolism categories (Table 2), to validate the microarray results for expression of these genes. Aconsistent agreement between the two methods was observed (Fig. 4). Together, these observations provide evidence that mtDNA mutations result in down-regulation of genes involved in energy metabolism in the cochlea of aged mitochondrial mutator mice, consistent with mitochondrial dysfunction and associated impairment of energy metabolism.
3.5. mtDNA mutations result in down-regulation of genes involved in cytoskeletal regulation
The EASE analysis revealed that six cytoskeletal regulation categories, including “cytoskeleton”, “extracellular matrix”, “collagen”, “structural molecule activity”, “myosin”, and “cell adhesion”, were significantly suppressed in D257A mice (Fisher Exact score < 0.05), and that 132 genes in these categories were significantly down-regulated as a result of age-related accumulation of mtDNA mutations (P < 0.02, FDR < 10%) (Table 1). Of the 132 genes identified, 39 genes were associated with the “cytoskeleton” category, including 5 myosin encoding genes: Myl1, Myh1, Myl2, Myh2, and Mylpf (Table 2). Myosins are a family of actin-based molecular motor proteins and play an important role in the auditory system (Gillespie, 2002). For example, the protein encoded by Myo7a is localized along the actin-rich stereocilia of hair cells in the cochlea (Gillespie, 2002; Wolfrum et al., 1998). Further, specific mutations in myo7a, myo15, and myo6 are associated with deafness (Kros et al., 2002; Melchionda et al., 2001). Twenty-five genes in the “extracellular matrix” category were significantly downregulated, including eight collagen encoding genes: Col3a1, Col1a2, Col5a2, Col4a1, Col17a1, Col15a1, and Col6a3 (Table 2). Type IX collagen is thought to play a key structural role in the tectorial membrane of the organ of Corti in cochlea (Suzuki et al., 2005). Specific mutations in Col1a1, Col4a3, Col4a4, Col4a5, and Col11a2 cause conductive hearing loss in humans (Chen et al., 2005; Gratton et al., 2005; Iliadou et al., 2006). The EASE analysis also revealed that 22 genes in the “cell adhesion” category were significantly down-regulated, including cadherin 13 (Cdh13) (Table 2), a member of a cadherin family that mediates cell–cell adhesion or cell-extracellular matrix interactions (Zheng et al., 2005). Specific mutations in Cdh23 and Pcdh15 cause hearing loss in mice (Noben-Trauth et al., 2003; Zheng et al., 2005). QRT-PCR analysis was conducted for Col17a1, to validate the microarray results for expression of these genes. A consistent agreement between the two methods was observed (Fig. 4). These results provide evidence that mtDNA mutations result in down-regulation of genes involved in cytoskeletal regulation in the cochlea of aged mitochondrial mutator mice, consistent with cytoskeletal dysfunction.
3.6. mtDNA mutations result in down-regulation of genes associated with hearing function
Gene expression profiling revealed that six genes involved in the “perception of sound” category were down-regulated, including procollagen, type I, alpha 1 (Col1a1) and otogelin (Otog) (P < 0.02, FDR < 10%) (Table 2). Otosclerosis, a bony dystrophy of the otic capsule caused by abnormal resorption and redeposition of bony tissue, is one of the common causes of hearing loss (Rodriguez et al., 2004). Hearing loss in this disease most commonly results from the fixation of the stapedial footplate to the oval window, which prevents ossicular vibration in response to sound (Rodriguez et al., 2004). Specific mutations in Col1a1 have been associated with this disorder (Rodriguez et al., 2004). The Otog gene encodes otogelin, an N-glycosylated protein, which is thought to play a key role in the deflection of the stereocilia bundles in the hair cells of the cochlea (Simmler et al., 2000). Specific mutations in this gene also cause deafness (Simmler et al., 2000).
In the cochlea, ion transport is needed to maintain proper ionic balance essential for hearing function (Jespersen et al., 2005). The EASE analysis revealed that 12 genes in the “ion transport” category were down-regulated, including gamma-aminobutyric acid (GABA-A) receptor, subunit alpha 1 (Gabra1), glutamate receptor, ionotropic, AMPA2 (alpha 2) (Gria2), and potassium voltage-gated channel, shaker-related, subfamily, member 6 (Kcna6) (P < 0.02, FDR < 10%, Fisher Exact score = 0.79) (Table 2). The protein encoded by Gabra1 is a receptor subunit for gamma-aminobutyric acid (GABA), a major inhibitory neurotransmitter in mammalian brain (Ma et al., 2005; Rosmond et al., 2002). Gria2 encodes a glutamate receptor subunit that is involved in glutamatergic transmission (Mead and Stephens, 2003). QRT-PCR analysis was conducted for Col1a1, to validate the microarray results for expression of these genes. A consistent agreement between the two methods was observed (Fig. 4). These observations were consistent with the ABR findings, and together provide evidence that mtDNA mutations result in down-regulation of genes involved in hearing function in the cochlea of aged mitochondrial mutator mice.
3.7. TUNEL staining and immunostaining
Apoptosis results in nuclear DNA fragmentation (Rich et al., 2000). To confirm whether levels of apoptosis are increased in the cochleae of aged D257A mice, we performed the TUNEL staining to examine cochleae from 2-and 9-month-old WT and D257A mice. We selected spiral ganglion cells in the apical region for evaluation of TUNEL staining because spiral ganglion cells were severely degenerated in the basal region of 9-month-old D257A mice. The mean percentage of TUNEL-positive cells in 9-month-old D257A mice (4.1%) was significantly elevated compared to those of 2-month-old WT (0.8%) and D257A (1.0%), and 9-month-old WT mice (1.5%) (P < 0.05, n = 5) (Fig. 5A–E).
Fig. 5.
Apoptosis detection by TUNEL staining. Detection and quantification of DNA fragmentation in the cochlear apical turn from 2- and 9-month-oldWTand D257A (A–D) mice. Arrows indicate TUNEL-positive staining cells. The cochleae from 9-month-old D257A mice displayed significantly more TUNEL-positive cells relative to those of age-matched WT (E) (*P < 0.05, n = 5). Error bars represent S.E.M. Scale bar = 25µm.
An additional indicator of apoptotic cell death is the activation of caspases (Hengartner, 2000; Yuan and Yankner, 2000). To examine whether levels of activated caspase-3 are increased in the cochleae of aged D257A mice, we examined cochleae of 2- and 9-month-old WT and D257A mice by immunostaining using an antibody for cleaved caspase-3. We also selected spiral ganglion cells in the apical region for this study. The mean percentage of cleaved caspase-3-positive cells in 9-month-old D257A mice (3.7%) was significantly elevated compared to that seen in 2-monthold WT (0.5%) and D257A (0.7%), and 9-month-old WT mice (1.2%) (P < 0.05, n=5) (Fig. 6A–E). These results were consistent with the microarray findings, and together provide evidence that mtDNA mutations result in induction of apoptosis in the cochlea of aged mitochondrial mutator mice.
Fig. 6.
Apoptosis detection by active caspase-3 staining. Detection and quantification of cleaved caspase-3 in the cochlear apical turn from 2- and 9- month-oldWTand D257A (A–D) mice by immunostaining. Arrows indicate cleaved caspase-3-positive staining cells. The cochleae from 9-month-old D257A mice displayed significantly more cleaved caspase-3-positive cells relative to those of age-matched WT (E) (*P < 0.05, n = 5). Error bars represent S.E.M. Scale bar = 25µm.
4. Discussion
The data presented here provide the first comprehensive gene expression analysis of a tissue in aged mitochondrial mutator mice and reveal previously unreported pathways associated with age-related hearing loss that may provide insights into pathogenic mechanisms. It has been postulated that mitochondrial function declines with aging partly due to oxidative damage to mtDNA, and that the resulting mitochondrial dysfunction leads to impairment of energy metabolism. Indeed, mitochondrial ATP content and production has been shown to decline with aging in the rat muscle by ~50% (Dirks et al., 2006). Mitochondrial ATP production also declines with advancing aging in human skeletal muscle (Short et al., 2005). Furthermore, mitochondrial mutator mice display a decline in respiratory enzyme activities and mitochondrial ATP production rates with aging in the heart, suggesting a role of mtDNA mutations as a cause of energy metabolism impairment (Trifunovic et al., 2004). In the present study, we observed transcriptional evidence for a decline in carbohydrate metabolism and mitochondrial function in the cochlea of aged mitochondrial mutator mice, suggesting that accumulation of mtDNA mutations leads to energy metabolism impairment. We have shown previously that the frequency of mtDNA mutations in the cytochrome b gene region from 5-month-old D257A mice was ~3–8 times that in the age-matched WT mice for heart, liver, and duodenum (Kujoth et al., 2005). An independent group has also reported that the frequency of somatic mtDNA point mutations in the cytochrome b gene region from 8-week-old mitochondrial mutator mice was approximately ~5 times higher than that of the age-matched wild-type mice for brain, and that further increases at 25 weeks were not significant (Trifunovic et al., 2004). These reports suggest that although a substantial load of mtDNA mutations may be present by 2 months of age in the cochlea of D257A mice, it takes several more months to develop age-related hearing loss in these animals. We did not measure the frequency of mtDNA mutations in the cochlear tissue of D257A mice, but similar to all other tissues examined in this model, it is likely that these mice have a substantial burden of mtDNA mutations in this tissue. With aging, this may lead to mitochondrial dysfunction, resulting in impairment of energy metabolism.
A progressive decline in mitochondrial function and associated energy depletion in aging muscle have been associated with sarcopenia, an age-related loss of muscle mass and strength (Dirks et al., 2006). It has been postulated that apoptosis induced by mitochondrial dysfunction and associated impairment of energy metabolism in aged muscle tissue may contribute to the development of sarcopenia (Dirks et al., 2006). Indeed, impairment of respiratory function is associated with induction of apoptosis (Dirks et al., 2006; Kujoth et al., 2005). We have previously reported that aged mitochondrial mutator mice display an increase in the levels of apoptotic markers in liver, testes, thymus, intestine, and muscle (Kujoth et al., 2005). Furthermore, Parkinson’s disease disease, one of the well-known age-related neuronal diseases, has also been associated with apoptotic death of neurons and mitochondrial mutations (Mattson, 2000). Interestingly, caloric restriction, known to retard aging and age-related diseases, has shown to suppress apoptotic cell death in the mouse cochlea and prevents age-related hearing loss (Someya et al., 2007), implicating apoptosis in the development of age-related hearing loss. In the present study, we observed up-regulation of over 50 genes associated with apoptotic pathways and a marked increase in the levels of apoptotic markers in the aging cochlea. Because D257A mice have a substantial burden of mtDNA mutations in postmitotic tissues, it is possible that these mtDNA mutations lead to mitochondrial dysfunction and associated impairment of energy metabolism, resulting in activation of apoptosis. Together, our observations and these cited reports suggest that apoptosis may play a major role in the pathogenesis of age-related hearing loss in mammals, and that mtDNA mutations is likely to induce apoptotic cell death in the mammalian cochlea.
Structural proteins such as myosins, collagens, and cadherins play a key role in maintaining the structural integrity and motor activity of cochlear cells and mediating the cell-to-cell interaction essential for auditory function (Gillespie, 2002; Suzuki et al., 2005; Zheng et al., 2005). Indeed, defects or mutations in a number of myosin, collagen, and cadherin genes have been associated with hearing loss (Chen et al., 2005; Gratton et al., 2005; Iliadou et al., 2006; Kros et al., 2002; Melchionda et al., 2001; Noben-Trauth et al., 2003; Zheng et al., 2005). Of these mutations, Cdh23 mutations have been linked to age-related hearing loss in mice and humans (Noben-Trauth et al., 2003). Cadherin 23 is specifically expressed in the stereocillia of the hair cells in the cochlea and plays a key role in perception of sound (Noben-Trauth et al., 2003). The C57BL/6 strain, a mouse homozygous for the Cdh23 mutation, has increased susceptibility to early onset of presbycusis (Zheng and Johnson, 2001), indicating that this modifier gene plays a key role in its pathogenesis. It is possible that mtDNA mutations may lead to accelerating cochlear degeneration due to the Cdh23 mutations in the C57BL/6 background. Furthermore, it is likely that several other modifier genes contribute to presbycusis and our microarray observations raise the possibility that these unknown modifiers might include genes involved in cytoskeletal regulation identified herein. We note that a dramatic down-regulation of the expression of Myh2 (fold change = −136.0), Lmod2 (−38.2), and Col17a1 (−14.0) genes was observed, implicating these genes as potential candidates for modifier or marker genes of age-related hearing loss. We also observed severe loss of spiral ganglion neurons and down-regulation of over one hundred genes involved in cytoskeletal regulation in D257A mice. Because impaired energy metabolism can lead to apoptosis, it is possible that loss of critical cells such as spiral ganglion neurons due to apoptosis may cause damage to the structural proteins of cochlea, resulting in cytoskeletal dysfunction. Together, our observations suggest that cytoskeletal dysfunction may be a key feature of age-related hearing loss.
Our findings suggest a previously unreported model of how mtDNA mutations impact cochlear function (Fig. 7). In this model, Polg infidelity causes accumulation of mtDNA mutations, leading to mitochondrial dysfunction and associated impairment of energy metabolism. The energy depletion in turn induces apoptosis, leading to loss of critical cochlear cells such as hair cells and spiral ganglion neurons. These processes eventually cause loss of cochlear function, resulting in age-related hearing loss. The relevance of this model to human presbycusis, mitochondrial diseases, and aging in general is worthy of further exploration.
Fig. 7.
Model of how mtDNA mutations impact cochlear function. In this model, aPolg mutation causes accumulation ofmtDNAmutations, leading to mitochondrial dysfunction and associated impairment of energy metabolism. The energy depletion in turn induces apoptosis, leading to loss of critical cochlear cells such as hair cells and spiral ganglion neurons. These processes eventually cause loss of cochlear function, resulting in age-related hearing loss.
Acknowledgements
We thank S. Kinoshita and Y. Kurasawa for histological processing. This research was supported by NIH grants AG021905 (T.A.P.), AG18922 (R.W.), MEXT grant (the National Project on Protein Structural and Functional Analyses) (M.T.), MHLW grants 15110201 (T.Y), 13470357 (T.Y. and M.T.).
Footnotes
Conflict of interest
The authors declare no actual or potential conflicts of interest.
Appendix A. Supplementary data
Supplementary information associated with this article can be found, in the online version, at doi:10.1016/j.neurobiolaging.2007.01.014.
References
- Affymetrix. Expression Analysis Technical Manual. Santa Clara, CA: Affymetrix; 2004. [Google Scholar]
- Amiry-Moghaddam M, Lindland H, Zelenin S, Roberg BA, Gundersen BB, Petersen P, Rinvik E, Torgner IA, Ottersen OP. Brain mitochondria contain aquaporin water channels: evidence for the expression of a short aqp9 isoform in the inner mitochondrial membrane. FASEB J. 2005;19:1459–1467. doi: 10.1096/fj.04-3515com. [DOI] [PubMed] [Google Scholar]
- Anderson S, Bankier AT, Barrell BG, de Bruijn MH, Coulson AR, Drouin J, Eperon IC, Nierlich DP, Roe BA, Sanger F, Schreier PH, Smith AJ, Staden R, Young IG. Sequence and organization of the human mitochondrial genome. Nature. 1981;290:457–465. doi: 10.1038/290457a0. [DOI] [PubMed] [Google Scholar]
- Aratani Y, Koyama H, Nyui S, Suzuki K, Kura F, Maeda N. Severe impairment in early host defense against candida albicans in mice deficient in myeloperoxidase. Infect. Immun. 1999;67:1828–1836. doi: 10.1128/iai.67.4.1828-1836.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Balaban RS, Nemoto S, Finkel T. Mitochondria, oxidants, and aging. Cell. 2005;120:483–495. doi: 10.1016/j.cell.2005.02.001. [DOI] [PubMed] [Google Scholar]
- Chen W, Kahrizi K, Meyer NC, Riazalhosseini Y, Van Camp G, Najmabadi H, Smith RJ. Mutation of col11a2 causes auto-somal recessive non-syndromic hearing loss at the dfnb53 locus. J. Med. Genet. 2005;42:e61. doi: 10.1136/jmg.2005.032615. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cheng EH, Wei MC, Weiler S, Flavell RA, Mak TW, Lindsten T, Korsmeyer SJ. Bcl-2, bcl-x(l) sequester bh3 domain-only molecules preventing bax- and bak-mediated mitochondrial apoptosis. Mol. Cell. 2001;8:705–711. doi: 10.1016/s1097-2765(01)00320-3. [DOI] [PubMed] [Google Scholar]
- Chinnery PF, Elliott C, Green GR, Rees A, Coulthard A, Turnbull DM, Griffiths TD. The spectrum of hearing loss due to mitochondrial DNA defects. Brain. 2000;123((Pt 1)):82–92. doi: 10.1093/brain/123.1.82. [DOI] [PubMed] [Google Scholar]
- Chwieralski CE, Welte T, Buhling F. Cathepsin-regulated apoptosis. Apoptosis. 2006;11:143–149. doi: 10.1007/s10495-006-3486-y. [DOI] [PubMed] [Google Scholar]
- Corral-Debrinski M, Horton T, Lott MT, Shoffner JM, Beal MF, Wallace DC. Mitochondrial DNA deletions in human brain: regional variability and increase with advanced age. Nat. Genet. 1992;2:324–329. doi: 10.1038/ng1292-324. [DOI] [PubMed] [Google Scholar]
- Dennis G, Jr, Sherman BT, Hosack DA, Yang J, Gao W, Lane HC, Lempicki RA. David: Database for annotation, visualization, and integrated discovery. Genome Biol. 2003;4:P3. [PubMed] [Google Scholar]
- Dirks AJ, Hofer T, Marzetti E, Pahor M, Leeuwenburgh C. Mitochondrial DNA mutations, energy metabolism and apoptosis in aging muscle. Ageing Res. Rev. 2006;5:179–195. doi: 10.1016/j.arr.2006.03.002. [DOI] [PubMed] [Google Scholar]
- Fischel-Ghodsian N. Mitochondrial deafness. Ear Hear. 2003;24:303–313. doi: 10.1097/01.AUD.0000079802.82344.B5. [DOI] [PubMed] [Google Scholar]
- Francis HW, Ryugo DK, Gorelikow MJ, Prosen CA, May BJ. The functional age of hearing loss in a mouse model of presbycusis. II. Neuroanatomical correlates. Hear Res. 2003;183:29–36. doi: 10.1016/s0378-5955(03)00212-0. [DOI] [PubMed] [Google Scholar]
- Gao G, Westling J, Thompson VP, Howell TD, Gottschall PE, Sandy JD. Activation of the proteolytic activity of adamts4 (aggrecanase-1) by c-terminal truncation. J. Biol. Chem. 2002;277:11034–11041. doi: 10.1074/jbc.M107443200. [DOI] [PubMed] [Google Scholar]
- Gary DS, Mattson MP. Pten regulates akt kinase activity in hippocampal neurons and increases their sensitivity to glutamate and apoptosis. Neuromol. Med. 2002;2:261–269. doi: 10.1385/NMM:2:3:261. [DOI] [PubMed] [Google Scholar]
- Gillespie PG. Myosin-viia and transduction channel tension. Nat. Neurosci. 2002;5:3–4. doi: 10.1038/nn0102-3. [DOI] [PubMed] [Google Scholar]
- Gratton MA, Rao VH, Meehan DT, Askew C, Cosgrove D. Matrix metalloproteinase dysregulation in the stria vascularis of mice with alport syndrome: implications for capillary basement membrane pathology. Am. J. Pathol. 2005;166:1465–1474. doi: 10.1016/S0002-9440(10)62363-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hengartner MO. The biochemistry of apoptosis. Nature. 2000;407:770–776. doi: 10.1038/35037710. [DOI] [PubMed] [Google Scholar]
- Hosack DA, Dennis G, Jr, Sherman BT, Lane HC, Lempicki RA. Identifying biological themes within lists of genes with ease. Genome Biol. 2003;4:R70. doi: 10.1186/gb-2003-4-10-r70. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Iliadou V, Van Den Bogaert K, Eleftheriades N, Aperis G, Vanderstraeten K, Fransen E, Thys M, Grigoriadou M, Pampanos A, Economides J, Iliades T, Van Camp G, Petersen MB. Monogenic nonsyndromic otosclerosis: audiological and linkage analysis in a large greek pedigree. Int. J. Pediatr. Otorhinolaryngol. 2006;70:631–637. doi: 10.1016/j.ijporl.2005.08.012. [DOI] [PubMed] [Google Scholar]
- Jespersen T, Grunnet M, Olesen SP. The kcnq1 potassium channel: from gene to physiological function. Physiology (Bethesda) 2005;20:408–416. doi: 10.1152/physiol.00031.2005. [DOI] [PubMed] [Google Scholar]
- Keithley EM, Canto C, Zheng QY, Fischel-Ghodsian N, Johnson KR. Age-related hearing loss and the ahl locus in mice. Hear Res. 2004;188:21–28. doi: 10.1016/S0378-5955(03)00365-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kros CJ, Marcotti W, van Netten SM, Self TJ, Libby RT, Brown SD, Richardson GP, Steel KP. Reduced climbing and increased slipping adaptation in cochlear hair cells of mice with myo7a mutations. Nat. Neurosci. 2002;5:41–47. doi: 10.1038/nn784. [DOI] [PubMed] [Google Scholar]
- Kujoth GC, Hiona A, Pugh TD, Someya S, Panzer K, Wohlgemuth SE, Hofer T, Seo AY, Sullivan R, Jobling WA, Morrow JD, Van Remmen H, Sedivy JM, Yamasoba T, Tanokura M, Weindruch R, Leeuwenburgh C, Prolla TA. Mitochondrial DNA mutations, oxidative stress, and apoptosis in mammalian aging. Science. 2005;309:481–484. doi: 10.1126/science.1112125. [DOI] [PubMed] [Google Scholar]
- Lee CK, Klopp RG, Weindruch R, Prolla TA. Gene expression profile of aging and its retardation by caloric restriction. Science. 1999;285:1390–1393. doi: 10.1126/science.285.5432.1390. [DOI] [PubMed] [Google Scholar]
- Levin S, Almo SC, Satir BH. Functional diversity of the phosphoglucomutase superfamily: structural implications. Protein Eng. 1999;12:737–746. doi: 10.1093/protein/12.9.737. [DOI] [PubMed] [Google Scholar]
- Ma DQ, Whitehead PL, Menold MM, Martin ER, Ashley-Koch AE, Mei H, Ritchie MD, Delong GR, Abramson RK, Wright HH, Cuccaro ML, Hussman JP, Gilbert JR, Pericak-Vance MA. Identification of significant association and gene–gene interaction of gaba receptor subunit genes in autism. Am. J. Hum. Genet. 2005;77:377–388. doi: 10.1086/433195. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mancuso M, Filosto M, Bellan M, Liguori R, Montagna P, Baruzzi A, DiMauro S, Carelli V. Polg mutations causing ophthalmoplegia, sensorimotor polyneuropathy, ataxia, and deafness. Neurology. 2004;62:316–318. doi: 10.1212/wnl.62.2.316. [DOI] [PubMed] [Google Scholar]
- Mattson MP. Apoptosis in neurodegenerative disorders. Nat. Rev. Mol. Cell Biol. 2000;1:120–129. doi: 10.1038/35040009. [DOI] [PubMed] [Google Scholar]
- Mead AN, Stephens DN. Involvement of ampa receptor glur2 subunits in stimulus-reward learning: Evidence from glutamate receptor gria2 knock-out mice. J. Neurosci. 2003;23:9500–9507. doi: 10.1523/JNEUROSCI.23-29-09500.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Melchionda S, Ahituv N, Bisceglia L, Sobe T, Glaser F, Rabionet R, Arbones ML, Notarangelo A, Di Iorio E, Carella M, Zelante L, Estivill X, Avraham KB, Gasparini P. Myo6, the human homologue of the gene responsible for deafness in snell’s waltzer mice, is mutated in autosomal dominant nonsyndromic hearing loss. Am. J. Hum. Genet. 2001;69:635–640. doi: 10.1086/323156. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Michikawa Y, Mazzucchelli F, Bresolin N, Scarlato G, Attardi G. Aging-dependent large accumulation of point mutations in the human mtdna control region for replication. Science. 1999;286:774–779. doi: 10.1126/science.286.5440.774. [DOI] [PubMed] [Google Scholar]
- Murashov AK, Talebian S, Wolgemuth DJ. Role of heat shock protein hsp25 in the response of the orofacial nuclei motor system to physiological stress. Brain. Res. Mol. Brain Res. 1998;63:14–24. doi: 10.1016/s0169-328x(98)00258-7. [DOI] [PubMed] [Google Scholar]
- Nguyen KV, Ostergaard E, Ravn SH, Balslev T, Danielsen ER, Vardag A, McKiernan PJ, Gray G, Naviaux RK. Polg mutations in alpers syndrome. Neurology. 2005;65:1493–1495. doi: 10.1212/01.wnl.0000182814.55361.70. [DOI] [PubMed] [Google Scholar]
- Noben-Trauth K, Zheng QY, Johnson KR. Association of cadherin 23 with polygenic inheritance and genetic modification of sensorineural hearing loss. Nat. Genet. 2003;35:21–23. doi: 10.1038/ng1226. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Orii KE, Aoyama T, Wakui K, Fukushima Y, Miyajima H, Yamaguchi S, Orii T, Kondo N, Hashimoto T. Genomic and mutational analysis of the mitochondrial trifunctional protein beta-subunit (hadhb) gene in patients with trifunctional protein deficiency. Hum. Mol. Genet. 1997;6:1215–1224. doi: 10.1093/hmg/6.8.1215. [DOI] [PubMed] [Google Scholar]
- Pickles JO. Mutation in mitochondrial DNA as a cause of presbyacusis. Audiol. Neurootol. 2004;9:23–33. doi: 10.1159/000074184. [DOI] [PubMed] [Google Scholar]
- Putcha GV, Moulder KL, Golden JP, Bouillet P, Adams JA, Strasser A, Johnson EM. Induction of bim, a proapoptotic bh3-only bcl-2 family member, is critical for neuronal apoptosis. Neuron. 2001;29:615–628. doi: 10.1016/s0896-6273(01)00238-0. [DOI] [PubMed] [Google Scholar]
- Ren J, Shi M, Liu R, Yang QH, Johnson T, Skarnes WC, Du C. The birc6 (bruce) gene regulates p53 and the mitochondrial pathway of apoptosis and is essential for mouse embryonic development. Proc. Natl. Acad. Sci. U.S.A. 2005;102:565–570. doi: 10.1073/pnas.0408744102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rich T, Allen RL, Wyllie AH. Defying death after DNA damage. Nature. 2000;407:777–783. doi: 10.1038/35037717. [DOI] [PubMed] [Google Scholar]
- Rodriguez L, Rodriguez S, Hermida J, Frade C, Sande E, Visedo G, Martin C, Zapata C. Proposed association between the col1a1 and col1a2 genes and otosclerosis is not supported by a case-control study in spain. Am. J. Med. Genet. A. 2004;128:19–22. doi: 10.1002/ajmg.a.30074. [DOI] [PubMed] [Google Scholar]
- Rosmond R, Bouchard C, Bjorntorp P. Allelic variants in the gaba(a)alpha6 receptor subunit gene (gabra6) is associated with abdominal obesity and cortisol secretion. Int. J. Obes. Relat. Metab. Disord. 2002;26:938–941. doi: 10.1038/sj.ijo.0802022. [DOI] [PubMed] [Google Scholar]
- Sakamoto T, Kondo K, Yamasoba T, Suzuki M, Sugasawa M, Kaga K. Overexpression of erbb-2 protein in human middle ear cholesteatomas. Laryngoscope. 2004;114:1988–1991. doi: 10.1097/01.mlg.0000147934.21638.d8. [DOI] [PubMed] [Google Scholar]
- Sawada S, Takeda T, Kitano H, Takeuchi S, Okada T, Ando M, Suzuki M, Kakigi A. Aquaporin-1 (aqp1) is expressed in the stria vascularis of rat cochlea. Hear Res. 2003;181:15–19. doi: 10.1016/s0378-5955(03)00131-x. [DOI] [PubMed] [Google Scholar]
- Schoenebeck B, Bader V, Zhu XR, Schmitz B, Lubbert H, Stichel CC. Sgk1, a cell survival response in neurodegenerative diseases. Mol. Cell Neurosci. 2005;30:249–264. doi: 10.1016/j.mcn.2005.07.017. [DOI] [PubMed] [Google Scholar]
- Seidman MD, Ahmad N, Bai U. Molecular mechanisms of age-related hearing loss. Ageing Res. Rev. 2002;1:331–343. doi: 10.1016/s1568-1637(02)00004-1. [DOI] [PubMed] [Google Scholar]
- Short KR, Bigelow ML, Kahl J, Singh R, Coenen-Schimke J, Raghavakaimal S, Nair KS. Decline in skeletal muscle mitochondrial function with aging in humans. Proc. Natl. Acad. Sci. U.S.A. 2005;102:5618–5623. doi: 10.1073/pnas.0501559102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Simmler MC, Cohen-Salmon M, El-Amraoui A, Guillaud L, Benichou JC, Petit C, Panthier JJ. Targeted disruption of otog results in deafness and severe imbalance. Nat. Genet. 2000;24:139–143. doi: 10.1038/72793. [DOI] [PubMed] [Google Scholar]
- Someya S, Yamasoba T, Weindruch R, Prolla TA, Tanokura M. Caloric restriction suppresses apoptotic cell death in the mammalian cochlea and leads to prevention of presbycusis. Neurobiol. Aging. 2007;28:1613–1622. doi: 10.1016/j.neurobiolaging.2006.06.024. [DOI] [PubMed] [Google Scholar]
- Suzuki N, Asamura K, Kikuchi Y, Takumi Y, Abe S, Imamura Y, Hayashi T, Aszodi A, Fassler R, Usami S. Type ix collagen knock-out mouse shows progressive hearing loss. Neurosci. Res. 2005;51:293–298. doi: 10.1016/j.neures.2004.12.001. [DOI] [PubMed] [Google Scholar]
- Trifunovic A, Wredenberg A, Falkenberg M, Spelbrink JN, Rovio AT, Bruder CE, Bohlooly YM, Gidlof S, Oldfors A, Wibom R, Tornell J, Jacobs HT, Larsson NG. Premature ageing in mice expressing defective mitochondrial DNA polymerase. Nature. 2004;429:417–423. doi: 10.1038/nature02517. [DOI] [PubMed] [Google Scholar]
- Trioulier Y, Torch S, Blot B, Cristina N, Chatellard-Causse C, Verna JM, Sadoul R. Alix, a protein regulating endosomal trafficking, is involved in neuronal death. J. Biol. Chem. 2004;279:2046–2052. doi: 10.1074/jbc.M309243200. [DOI] [PubMed] [Google Scholar]
- Tusher VG, Tibshirani R, Chu G. Significance analysis of microar-rays applied to the ionizing radiation response. Proc. Natl. Acad. Sci. U.S.A. 2001;98:5116–5121. doi: 10.1073/pnas.091062498. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wolfrum U, Liu X, Schmitt A, Udovichenko IP, Williams DS. Myosin viia as a common component of cilia and microvilli. Cell Motil. Cytoskeleton. 1998;40:261–271. doi: 10.1002/(SICI)1097-0169(1998)40:3<261::AID-CM5>3.0.CO;2-G. [DOI] [PubMed] [Google Scholar]
- Wuyts W, Reyniers E, Ceuterick C, Storm K, de Barsy T, Martin JJ. Myopathy and phosphorylase kinase deficiency caused by a mutation in the phka1 gene. Am. J. Med. Genet. A. 2005;133:82–84. doi: 10.1002/ajmg.a.30517. [DOI] [PubMed] [Google Scholar]
- Yin XM. Bid, a critical mediator for apoptosis induced by the activation of fas/tnf-r1 death receptors in hepatocytes. J. Mol. Med. 2000;78:203–211. doi: 10.1007/s001090000099. [DOI] [PubMed] [Google Scholar]
- Yuan J, Yankner BA. Apoptosis in the nervous system. Nature. 2000;407:802–809. doi: 10.1038/35037739. [DOI] [PubMed] [Google Scholar]
- Zheng QY, Johnson KR. Hearing loss associated with the modifier of deaf waddler (mdfw) locus corresponds with age-related hearing loss in 12 inbred strains of mice. Hear Res. 2001;154:45–53. doi: 10.1016/s0378-5955(01)00215-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zheng QY, Yan D, Ouyang XM, Du LL, Yu H, Chang B, Johnson KR, Liu XZ. Digenic inheritance of deafness caused by mutations in genes encoding cadherin 23 and protocadherin 15 in mice and humans. Hum. Mol. Genet. 2005;14:103–111. doi: 10.1093/hmg/ddi010. [DOI] [PMC free article] [PubMed] [Google Scholar]







